Absolute Bioavailability of Isoflavones from Soy Protein Isolate

Mar 15, 2010 - Jefferson, Arkansas 72079 ... supplements for their putative health benefits; however, evidence for potential adverse effects has been ...
0 downloads 0 Views 912KB Size
J. Agric. Food Chem. 2010, 58, 4529–4536

4529

DOI:10.1021/jf9039843

Absolute Bioavailability of Isoflavones from Soy Protein Isolate-Containing Food in Female Balb/c Mice JUAN E. ANDRADE,† NATHAN C. TWADDLE,‡ WILLIAM G. HELFERICH,*,† AND DANIEL R. DOERGE*,‡ † Department of Food Science and Human Nutrition, University of Illinois, Urbana-Champaign, Illinois 61801, and ‡National Center for Toxicological Research, U.S. Food and Drug Administration, Jefferson, Arkansas 72079

Soy isoflavones, genistein and daidzein, are widely consumed in soy-based foods and dietary supplements for their putative health benefits; however, evidence for potential adverse effects has been obtained from experimental animal studies. An important prerequisite for understanding the pharmacodynamics of isoflavones is better information about pharmacokinetics and bioavailability. This study determined the bioavailability of genistein and daidzein in a mouse model by comparing plasma pharmacokinetics of their aglycone and conjugated forms following administration of identical doses (1.2 mg/kg genistein and 0.55 mg/kg daidzein) by either an intravenous injection (IV) or gavage of the aglycones in 90% aqueous solution vs a bolus administration of equimolar doses delivered in a food pellet prepared using commercial soy protein isolate (SPI) as the isoflavone source. The bioavailability of genistein and daidzein was equivalent for the gavage and dietary routes of administration despite the use of isoflavone aglycones in the former and SPIderived glucosides in the latter. While absorption of total isoflavones was nearly quantitative from both oral routes [>84% of areas under the curve (AUCs) for IV], presystemic and systemic phase II conjugation greatly attenuated internal exposures to the receptor-active aglycone isoflavones (9-14% for genistein and 29-34% for daidzein based on AUCs for IV). These results show that SPI is an efficient isoflavone delivery vehicle capable of providing significant proportions of the total dose into the circulation in the active aglycone form for distribution to receptor-bearing tissues and subsequent pharmacological effects that determine possible health benefits and/or risks. KEYWORDS: Isoflavone; mice; bioavailability; soy protein isolate; pharmacokinetics; genistein

*To whom correspondence should be addressed. (D.R.D.) Tel: 870543-7943. Fax: 870-543-7720. E-mail: [email protected]. (W.G.H.) Tel: 217-244-5414. Fax: 217-244-9522. E-mail: helferic@ uiuc.edu.

These observations made it imperative to better understand how the food matrix can affect important isoflavone pharmacokinetic factors in an animal model. In the current study, we chose to evaluate the effects of soy protein isolate (SPI) because it is a major commercial ingredient used in the manufacture of many modern soy-based foods and dietary supplements, including infant formula. For these reasons, the current study examined the absolute bioavailability of genistein and daidzein, as both aglycones and total conjugated forms, in the plasma of female Balb/c mice. This was achieved by comparing the plasma pharmacokinetics from intravenous (IV) and gavage administration of genistein and daidzein aglycones in solution with bolus feeding of equimolar doses of isoflavone glucosides in a SPIcontaining food matrix that is comparable to many popular soy foods, such as high-protein “power bars”. The experimental design chosen in this study also specifically addresses common methodological deficiencies found in many studies of the pharmacokinetics of soy isoflavones. First, low doses of isoflavone glucosides relevant to food intake were chosen to minimize complications due to nonlinear pharmacokinetics often encountered as dose increases, and second, both aglycone and total isoflavone concentrations were measured in plasma.

© 2010 American Chemical Society

Published on Web 03/15/2010

INTRODUCTION

The bioavailability of isoflavones (Scheme 1) from soy-based foods and dietary supplements is a critical component in understanding possible beneficial and adverse health effects in animal models and humans. There is general agreement that absorption in the rodent and human gut proceeds by rapid release of aglycones through the action of bacterial and intestinal β-glucosidases following ingestion of isoflavone glucosides present in whole soy foods (1-5) (Scheme 2). While there is evidence for the delivery of small amounts of intact isoflavone glucosides into the circulation in rodents (6, 7) and humans (2), efficient phase II conjugation within the enterocyte and secretion of glucuronides into the blood greatly limit the amount of biologically active aglycone available for distribution to the tissues (i.e., bioavailable). There is strong evidence in rodents for the secretion of isoflavone glucuronides into the bile and enterohepatic recirculation prior to elimination into both feces and urine (7-10).

pubs.acs.org/JAFC

4530

J. Agric. Food Chem., Vol. 58, No. 7, 2010

Scheme 1. Structures of Isoflavones, Their β-D-Glucosides, and Equol

Scheme 2. Metabolic Transformations of Dietary Isoflavones That Affect Bioavailability

To accomplish these goals, analytical methodology based on liquid chromatography/tandem mass spectrometry and isotope dilution were employed to ensure sensitivity adequate to quantify circulating aglycones, which are the receptor-binding forms, and their inactive conjugated forms. MATERIALS AND METHODS Chemicals and Reagents. High-performance liquid chromatography (HPLC)-grade methanol and acetonitrile and reagent grade hydrochloric acid, glacial acetic acid, and ammonium acetate were from Fisher Scientific (Fair Lawn, NJ). Isoflavone standards, daidzin, glycitin, genistein, daidzein, glycitein, genistein, and the internal standard, formononetin, were purchased from different commercial vendors, and their purity was determined and confirmed using liquid chromatographymass spectrometry (LC-MS), NMR, and LC-UV. Briefly, highly purified chemicals (CAS) were purchased as follows: genistin (529-59-9), glycitein (40957-83-3), and formononetin (485-72-3) from Indofine Chemical Co. (Hillsborough, NJ) and daidzin (552-66-9) and glycitin (40246-10-4) from Sigma-Aldrich Chemical (St. Louis, MO). Both genistein (446-72-0) and daidzein (486-66-8) were synthesized in the Helferich laboratory to >98% purity as previously described (11). Standard solutions at 1-2 mg/mL were prepared in dimethyl sulfoxide (DMSO; Sigma-Aldrich Chemical)

Andrade et al. and stored in darkness at -20 C. Working standard solutions were prepared on the day of the analysis by dilution of the stock solutions with aqueous methanol and mobile phase A. All solutions were filtered through a 0.45 μm Acrodisc syringe filter (PALL Co.; East Hills, NY) prior to reverse-phase HPLC separation. The purity of the isoflavone aglycones and glucosides stock solutions was checked with maximum absorption and known molar extinction coefficients as described in Nurmi et al. (12). Eighteen separate samples of commercial SPIs were purchased from different manufacturers and distributors (Table 1) within a 2 year period (2005-2007). A sample of SPI was prepared using mild alkali conditions as described by Wu et al. (13) from soy flour (Bakers soy flour, ADM, Decatur, IL). Analysis of Isoflavones in Foods. Soy products were analyzed at least in duplicate for isoflavone composition using mild acid hydrolysis conditions as previously described (14). Acid hydrolysis solution was prepared from absolute ethyl alcohol (200 proof, AAPER; Shelbyville, KY), double-deionized water, and concentrated HCl (12.1 M) in a 64:16:20 (v/v/v) ration, respectively, for a final HCl concentration of 2.4 M. Isoflavones were extracted in aqueous methanol, separated by reversephase HPLC with gradient elution, and quantified using coulometric detection against pure standards. Nonhydrolytic Analysis. Briefly, 100 mg of each sample (in duplicates) was placed in 15 mL centrifuge tubes. An aliquot (10 μL) of formononetin internal standard solution (2.5 mg/mL in DMSO) was spiked on top of the sample, before adding 10 mL of aqueous (80%) methanol. Samples were mixed until materials were suspended in solution and placed in an ultrasound water bath (Fisher Scientific) for 30 min. Upon sonication, samples were diluted (1:2; v/v) in separate borosilicate glass tubes with mobile phase A, filtered with 0.22 μm Acrodisc syringe filter (PALL Co.), and injected (10 μL) into a HPLC 100 μL sample loop. Acid Hydrolysis Analysis. Mild acid hydrolysis conditions were used to convert isoflavones from their malonyl and acetyl esters to their β-glucoside forms as the basis to calculate total glucoside and aglycone contents for each sample. Each sample (100 mg in duplicate) was accurately weighed into ∼70 mL tall PTFE screw-capped tubes. An aliquot (10 μL) of formononetin internal standard solution (2.5 mg/mL in DMSO) was spiked on top of the sample before addition of 10 mL of acid hydrolysis solution. A small stirring bar was added, and tubes were tightly capped. Samples were mixed until materials were suspended in solution. Samples were heated in a stirring-heating block unit controlled with a thermostat set 80 C for 1 (partial hydrolysis) or 5 h (complete hydrolysis). Upon termination, samples were diluted (1:3; v/v) in separate borosilicate glass tubes with mobile phase A, filtered through 0.22 μm Acrodisc syringe filters directly into 4 mL amber vials, and injected (10 μL) into a HPLC 100 μL sample loop. Chromatographic conditions were adapted from those described by Pe~nalvo et al. (14). Isoflavone determination was carried out using HPLC with the multichannel coulometric array detection system (ESA Inc.; Chelmsford, MA) consisting of two solvent pumps (model 582), autosampler model 717-Plus with cooling unit (Waters; Milford, MA), and a CoulArray detector (model 5600) with two analytical cells in series, each with four porous graphite electrodes set at 450, 490, 545, 580, 620, 690, 760, and 800 mV, in the same order. Analytes were separated using a Zorbax RX C18 column (150 mm  4.6 mm, 5 μm, 80 A˚) from MAC-MAD Analytical (Chadds Ford, PA), protected with a guard column packed with C18 material (Upchurch Sci. Inc., Oak Harbor, WA). Samples were kept, injected, and separated at room temperature (∼23 C). Gradient elution was used for complete separation of the analytes and consisted of two eluents: mobile phase A (MPA), 50 mM ammonium acetate buffer, pH 4.5, and MeOH (90:10 v/v), respectively, and MPB, 50 mM ammonium acetate buffer, pH 4.5, MeOH, and ACN (20:30:50 v/v/v), respectively. Gradient elution of isoflavones was carried out at 1 mL/min in the following pattern: 0-6 min at 15% MPB, increase to 18% MPB in 10 min, keep at 18% MPB for 4 min, increase to 25% MPB in 7 min, keep at 25% MPB for 4 min, increase to 38% MPB in 4 min, keep at 38% MPB for 5 min, increase to 42% MPB in 7 min, increase to 90% MPB in 7 min, keep at 90% MPB for 1 min, reduce to initial conditions in 4 min, and equilibrate for 8 min, for a total run time of 67 min including 8 min of equilibration. Controlled data acquisition and processing were performed using CoulArray software (version 2.0). Quantification from the sum of peak heights was performed using calibration curves generated from four-point

Article

J. Agric. Food Chem., Vol. 58, No. 7, 2010

4531

Table 1. Isoflavone Profile in Different SPIs SPI

mg/g

%

product name

daidzein

glycitein

genistein

total

daidzein

glycitein

genistein

Gluca

Aglyca

ADM 161b ADM 674b Ardex-Fb Ardexb Profam 648b Profam 781b Profam 873b Profam 892b Profam 982b soy flour (SF)b SPI from SFc Solpro 910d Solpro 950d SNE-Le SNE-NOe SNE-Oe Prolissef H0206 PTIg Protient 6300h protein drinki minimum maximum mean SD CV%

0.399 0.524 0.093 0.100 0.101 0.100 0.325 0.290 0.106 0.491 0.522 0.288 0.311 0.106 0.067 0.355 0.203 0.849 0.086 0.619 0.067 0.849 0.297 0.218 73.3

0.092 0.161 0.039 0.048 0.035 0.051 0.083 0.075 0.037 0.150 0.118 0.091 0.090 0.026 0.030 0.061 0.045 0.174 0.020 0.084 0.020 0.174 0.075 0.046 60.7

0.672 0.677 0.288 0.437 0.251 0.309 0.820 0.607 0.312 0.734 1.171 0.902 0.972 0.195 0.162 1.243 0.399 1.260 0.236 1.159 0.162 1.260 0.640 0.377 58.9

1.162 1.362 0.420 0.585 0.387 0.459 1.229 0.972 0.456 1.375 1.812 1.281 1.373 0.328 0.259 1.658 0.647 2.282 0.341 1.861 0.259 2.282 1.013 0.610 60.3

34.3 38.5 22.2 17.1 26.1 21.7 26.5 29.8 23.3 35.7 28.8 22.5 22.6 32.4 26.0 21.4 31.3 37.2 25.1 33.3 17.1 38.5 27.8 6.0 21.7

7.9 11.8 9.2 8.2 9.1 11.0 6.8 7.7 8.2 10.9 6.5 7.1 6.6 7.9 11.4 3.7 7.0 7.6 5.7 4.5 3.7 11.8 7.9 2.2 27.5

57.8 49.7 68.5 74.7 64.7 67.3 66.8 62.5 68.5 53.4 64.6 70.4 70.8 59.6 62.6 75.0 61.7 55.2 69.2 62.3 49.7 75.0 64.3 6.8 10.5

96.8 96.4 79.2 83.2 65.2 78.4 86.9 90.3 80.1 98.2 97.8 89.1 89.7 23.0 92.6 81.8 18.9 80.6 90.4 92.8 18.9 98.2 80.6 22.0 27.3

3.2 3.6 20.8 16.8 34.8 21.6 13.1 9.7 19.9 1.8 2.2 10.9 10.3 77.0 7.4 18.2 81.1 19.4 9.6 7.2 1.8 81.1 19.4 22.0 113.2

a Gluc, glucoside; Aglyc, aglycone. b From Archer Daniels Midland (ADM). c From Helferich Laboratory. d From Solbar, Ltd. (Ashdod, Israel). e SNE, Soy-N-Ergy, plus lecithin, nonorganic, or organic from American Health & Nutrition (Ann Arbor, MI). f From Cargill, Inc. (Minneapolis, MN). g From Protein Technologies International, Inc. (St. Louis, MO). h From Protient, Inc. (St. Paul, MN). i Iso-Rich Soy high protein drink where SPI is the main ingredient (>90%) from Jarrow Formulas (Los Angeles, CA).

serial dilutions of single standards run in triplicates. Unknown peaks were matched to standards on the basis of retention time ((3%) as well as response ratio between adjacent channels ((35%) and manually inspected to ensure correct assignment. Diet Formulation. American Institute of Nutrition 93 growth (AIN 93G) semipurified diet, with corn oil substituting soybean oil, was selected as a basal diet for control animals, which meets all of the nutritional requirements of mice (15, 16). SPI with its endogenous isoflavone content was selected to evaluate the pharmacokinetics of genistein and daidzein in mice after consumption of a suitable high-protein food. Chocolate-based pellets, with or without SPI, were prepared by grinding commercial chocolate cookies and adding 20 mg of SPI (20% by weight). Pellets weighed ∼100 mg ((2 mg) each and were compounded to a pie shape (diameter ∼6.8 mm and height ∼2.64 mm) using a specially designed stainless steel shaping device and a hammer. Pellets without SPI were used to train the mice to consume the entire meal as a bolus. The same SPI without isoflavones was used to prepare pellets for the control group. Isoflavones were removed from SPI after ethanol washing and were confirmed to have negligible amounts of isoflavones (data not shown). This high-protein food was formulated similarly to commercially available high-protein bar products. According to our calculations and using the U.S. Department of Agriculture Food Composition database (17) and available composition tables from vendors, each 100 mg pellet contained protein (17%), carbohydrates (59%), and fats (17%), which delivered ∼1.8 kJ/pellet representing ∼5.7% of the daily energy requirements for mice used in this study (15, 16). Animal Handling Procedures. Female Balb/c mice (16.7 ( 1.2 g) were purchased from the National Cancer Institute (Bethesda, MD) and delivered at 30 days of age. Animals were housed individually and kept on a 12 h dark/12 h light cycle. All experiments were approved by the Institutional Animal Care and Use Committee of the University of Illinois. Animals were changed to AIN 93G basal diets on the day after arrival. Mice were trained over a 2 week period to consume the chocolate-based pellets (without SPI) in a bolus manner (i.e., within a few minutes). At the end of the training period, animals were randomized by weight and assigned to either the control or the treatment group. Animals in the control group (n = 6) received pellets prepared with ethanol-washed SPI.

The treatment group consisted of 42 mice (i.e., six mice in each of seven time points) and received cookie snacks prepared with intact SPI each containing 0.009 ( 0.001 and 0.019 ( 0.003 mg of daidzein and genistein, respectively, delivering respective doses of 0.55 and 1.16 mg/kg bw. On the day of sample collection, body weights were recorded, and mice were appropriately dosed, and blood samples were collected by cardiac puncture at 30, 60, 120, 240, 480, 720, and 1440 min after the feeding period. Control animals were euthanized at 30 min after the consumption of control pellets. Plasma was prepared by collection of blood into ethylenediaminetetraacetic acid-containing microcentrifuge tubes for further centrifugation at 2000 rpm (30 min, 4 C). Samples were kept at -80 C until further analyses. Pharmacokinetic evaluations of intravenous and gavage administrations were also conducted using female Balb/c mice similarly obtained from the National Cancer Institute. Procedures involving care and handling of these mice were reviewed and approved by the NCTR Laboratory Animal Care and Use Committee. Body weights were 18.7 ( 0.87 g (n = 107) and did not significantly differ between control, gavage, and IV groups. Dosing solutions were prepared less than 24 h before administration by dissolving genistein (4.6 mg) and daidzein (2.1 mg) in 4 mL of DMSO and then diluted to 40 mL with water to make a clear solution. A dose of 1.2 mg/kg genistein and 0.55 mg/kg daidzein was administered by either IV injection into the tail vein or by gavage using a dosing solution containing 115 μg/mL genistein and 55 μg/mL daidzein with 10 μL administered for every g body weight. The same dosing solution was used for IV and gavage administrations. Following dosing, six mice were sacrificed at each of the time points (5, 15, 30, 60, 120, 240, 360, 480, and 1080 min for IV and gavage) by CO2 asphyxiation, and blood was collected immediately by cardiac puncture. Blood samples were placed in purple top tubes, plasma was produced by centrifugation, and samples were stored at -80 C until analyzed in a single batch. Quantification of Plasma Isoflavones. Plasma concentrations of total isoflavones were determined after complete enzymatic hydrolysis using a H. pomatia preparation containing glucuronidase, sulfatase, and glucosidase activities using a previously validated LC-ES/MS/MS method based on isotope dilution quantification of genistein, daidzein, and equol (18). Isoflavone aglycones were determined identically except without

4532

J. Agric. Food Chem., Vol. 58, No. 7, 2010

Andrade et al. Table 2. Pharmacokinetic Parameters for Total and Aglycone Genistein in Plasma Following Administration to Mice by Different Routesa IV parameters

total

dose (μmol/kg bw) 4.3 85 t1/2 elim (min)b AUC0-¥ (μmol min/L) 39 Vd (L/kg bw) 14 Cl (L/min/kg bw) 0.11 bioavailability % aglycone

gavage

aglycone 9.8 11 5.2 0.37

total

aglycone

4.3 82 34

14 1.1

0.89

0.094 3.2

28

SPI total

aglycone

4.3 74 33

57 1.6

0.86

0.14 4.9

a

Absolute bioavailability (fraction absorbed) was computed as the ratio of AUC for oral administration to that for intravenous (IV) administration. b t1/2 elim, elimination half-time; AUC0-¥, area under the time-concentration curve from zero to infinity; Vd, volume of distribution; Cl, total plasma clearance; and bioavailability, AUC oral/AUC IV.

Table 3. Pharmacokinetic Parameters for Total and Aglycone Daidzein in Plasma Following Administration to Mice by Different Routesa IV parameters

total

dose (μmol/g bw) 2.1 160 t1/2 elim (min)b AUC0-¥ (μmol min/L) 34 Vd (L/kg bw) 14 Cl (L/min/kg bw) 0.061 bioavailability % aglycone

Figure 1. Plasma concentration-time profiles for aglycone and total genistein and daidzein following intravenous administration. The top panel shows a plot of plasma concentrations for aglycone genistein (closed circles) and daidzein (open circles) vs time following an intravenous injection of 1.2 and 0.55 mg/kg bw, respectively. The bottom panel shows a plot of plasma concentrations for total (i.e., determined following complete enzymatic hydrolysis of conjugated forms) genistein (closed circles) and daidzein (open circles). enzymatic deconjugation. Method detection limits (s/n ratio of 3) for genistein, daidzein, and equol were approximately 0.005 μM for analysis of 10 μL of plasma (total isoflavones) and 0.001 μM for analysis of 50 μL of plasma (aglycone isoflavones). Intra- and interday precisions were in the range of 3-13% relative standard deviation (SD), and accuracy was in the range of 88-99% (18). Quality control analyses were also performed during every sample set and included the analysis of blank and spiked serum samples (including glucuronidase/sulfatase), blank injections, and injections of authentic standards. Pharmacokinetic Analysis. Model-independent pharmacokinetic analysis was performed using PK Solutions 2.0 software (Summit Research Services, Montrose, CO). Parameters reported include the following: t1/2 elim, elimination half-time; AUC0-¥, area under the timeconcentration curve from zero to infinity; Vd, volume of distribution; Cl, total plasma clearance; and bioavailability, AUC oral/AUC IV. Statistical Analysis. Data are shown as the means ( SDs. All statistical analyses were conducted using SAS software (SAS Institute Inc., Cary, NC). Group comparisons were made using the two-sided t test. All reported P values are two-tailed, and differences at P < 0.05 were considered statistically significant. RESULTS

In this study, we first determined the total quantity of isoflavones and their profile in multiple commercial SPIs and one

gavage

aglycone 10 9.3 3.3 0.22

total

aglycone

total

aglycone

2.1 200 42

60 3.1

2.6 190 41c

42 2.8c

1.2 27

SPI

0.34 7.4

1.2

0.29 6.7

a Absolute bioavailability (fraction absorbed) was computed as the ratio of AUC for oral administration to that for IV administration. b t1/2 elim, elimination half-time; AUC0-¥, area under the time-concentration curve from zero to infinity; Vd, volume of distribution; Cl, total plasma clearance; and bioavailability, AUC oral/AUC IV. c Daidzein AUCs were adjusted for the small difference in administered dose for the SPI treatment group.

SPI made in our laboratory from defatted soy flour. We found that both the content and the profile of isoflavones in SPIs, even in those from the same supplier, are variable with average aglycone equivalents in mg/g of SPI of 0.297 ( 0.218, 0.075 ( 0.046, and 0.640 ( 0.377 mg/g (mean ( SD) for daidzein, glycitein, and genistein, respectively (Table 1). Similarly, the variation in the profile of isoflavones, total glycosides vs total aglycones (80.6 ( 22.0 vs 19.4 ( 22.0%), was quite large with most of the SPIs, with the exception of two, containing isoflavones in their original glycoside form as is common with most unfermented soy foods (19). Nonetheless, the relative relationship among aglycone equivalents in all SPIs was more constant with averages of 27.8 ( 6.0, 7.9 ( 2.2, and 64.3 ( 6.8% (mean ( SD) for daidzein, glycitein, and genistein, respectively. Thus, although the quantity and profile of isoflavones were quite variable, the relative proportion among aglycones was more constant. Next, we selected a SPI product (Solpro 950) based on the amounts and profile of its constituent isoflavones to determine the dose to use for evaluating the pharmacokinetics of genistein and daidzein (1.2 mg/kg bw genistein and 0.55 mg/kg bw daidzein). IV administration of genistein and daidzein produced circulating levels of aglycone and total isoflavones (i.e., conjugates þ aglycones) that decreased rapidly (Figure 1 and Tables 2 and 3). Half-times for elimination of aglycones were quite fast (10 min) with slower elimination of the total isoflavones (85 and 160 min for genistein and daidzein, respectively). The greater Vd for aglycone genistein relative to daidzein is consistent with its greater lipophilicity and more extensive distribution into tissues. Significant systemic phase II metabolism was evident because even at the initial time point following IV administration (5 min), only

Article

Figure 2. Plasma concentration-time profiles for aglycone and total genistein and daidzein following gavage administration. The top panel shows a plot of plasma concentrations for aglycone genistein (closed circles) and daidzein (open circles) vs time following an gavage administration of 1.2 and 0.55 mg/kg bw, respectively. The bottom panel shows a plot of plasma concentrations for total (i.e., determined following complete enzymatic hydrolysis of conjugated forms) genistein (closed circles) and daidzein (open circles).

36-70% of genistein and 36-81% of daidzein were present as aglycones in individual mice. Over the whole exposure period following IV administration, the ratio of AUCs for aglycone/total was 28% for genistein and 27% for daidzein. The similar systemic phase II metabolism for these compounds as reflected by these AUC percentages is consistent with the comparable catalytic activity shown by human liver microsomes for genistein (kcat/ Km = 0.093 pmol/min/mg protein) vs daidzein (kcat/Km = 0.11 pmol/min/mg protein) (20). No clear evidence for enterohepatic recirculation (i.e., an apparent “bump” in plasma concentrations late in the profile) was provided by these data; however, the statistical limitations imposed by interanimal variability from the chosen experimental design of using multiple mice at each time point could mitigate this interpretation. Administration of identical doses of genistein and daidzein by gavage produced more complex plasma concentration-time profiles (Figure 2). The impact of presystemic phase II metabolism in the gut was evident because in individual mice only 3-13% of genistein and 10-15% of daidzein were present as aglycones at the initial sampling point. Over the whole exposure period, the ratio of AUCs for aglycone/total was 3.2% for genistein and 7.4% for daidzein. The changes in relative AUCs between aglycone and total isoflavones from either IV or oral

J. Agric. Food Chem., Vol. 58, No. 7, 2010

4533

Figure 3. Plasma concentration-time profiles for aglycone and total genistein and daidzein following bolus administration of a SPI-containing food pellet. The top panel shows a plot of plasma concentrations for aglycone genistein (closed circles) and daidzein (open circles) vs time following consumption of 1.2 and 0.55 mg/kg bw, respectively, from a SPIcontaining food pellet. The bottom panel shows a plot of plasma concentrations for total (i.e., determined following complete enzymatic hydrolysis of conjugated forms) genistein (closed circles) and daidzein (open circles).

administration suggest that presystemic (i.e., gut) metabolism of genistein (28 to 3.2%, respectively) is more extensive than for daidzein (27 to 7.4%, respectively). This conclusion is consistent with the preferential glucuronidation of genistein (kcat/Km = 56 pmol/min/mg protein) over daidzein (not detected, 86%), active phase II metabolism in the gut and liver significantly attenuate internal exposure to the biologically active aglycones (9-14% for genistein and 29-34% for daidzein). Nonetheless, these results show that SPI can efficiently deliver significant proportions of the total administered oral dose into the circulation in the active aglycone form for distribution to receptor-bearing tissues and subsequent pharmacological effects. This clearer definition of internal exposures to the active forms of isoflavones should be considered in discussions of putative beneficial and potentially adverse health effects following consumption of SPI-based foods from the constellation of such products currently available in the marketplace. LITERATURE CITED (1) Day, A. J.; Canada, F. J.; Diaz, J. C.; Kroon, P. A.; McLauchlan, R.; Faulds, C. B.; Plumb, G. W.; Morgan, M. R.; Williamson, G. Dietary flavonoid and isoflavone glycosides are hydrolysed by the lactase site of lactase phlorizin hydrolase. FEBS Lett. 2000, 468 (2-3), 166–170. (2) Hosoda, K.; Furuta, T.; Yokokawa, A.; Ogura, K.; Hiratsuka, A.; Ishii, K. Plasma profiling of intact isoflavone metabolites by highperformance liquid chromatography and mass spectrometric identification of flavone glycosides daidzin and genistin in human plasma after administration of kinako. Drug. Metab. Dispos. 2008, 36 (8), 1485–1495. (3) Hawksworth, G.; Drasar, B. S.; Hill, M. J. Intestinal bacteria and the hydrolysis of glycosidic bonds. J. Med. Microbiol. 1971, 4 (4), 451– 459. (4) Chang, Y. C.; Nair, M. G. Metabolism of daidzein and genistein by intestinal bacteria. J. Nat. Prod. 1995, 58 (12), 1892–1896. (5) Sfakianos, J.; Coward, L.; Kirk, M.; Barnes, S. Intestinal uptake and biliary excretion of the isoflavone genistein in rats. J. Nutr. 1997, 127 (7), 1260–1268.

J. Agric. Food Chem., Vol. 58, No. 7, 2010

4535

(6) Allred, C. D.; Twaddle, N. C.; Allred, K. F.; Goeppinger, T. S.; Churchwell, M. I.; Ju, Y. H.; Helferich, W. G.; Doerge, D. R. Soy processing affects metabolism and disposition of dietary isoflavones in ovariectomized BALB/c mice. J. Agric. Food Chem. 2005, 53 (22), 8542–8550. (7) Steensma, A.; Faassen-Peters, M. A.; Noteborn, H. P.; Rietjens, I. M. Bioavailability of genistein and its glycoside genistin as measured in the portal vein of freely moving unanesthetized rats. J. Agric. Food Chem. 2006, 54 (21), 8006–8012. (8) Coldham, N. G.; Zhang, A. Q.; Key, P.; Sauer, M. J. Absolute bioavailability of [14C] genistein in the rat; plasma pharmacokinetics of parent compound, genistein glucuronide and total radioactivity. Eur. J. Drug Metab. Pharmacokinet. 2002, 27 (4), 249–258. (9) Moon, Y. J.; Sagawa, K.; Frederick, K.; Zhang, S.; Morris, M. E. Pharmacokinetics and bioavailability of the isoflavone biochanin A in rats. AAPS J. 2006, 8 (3), E433–E442. (10) Sepehr, E.; Cooke, G. M.; Robertson, P.; Gilani, G. S. Effect of glycosidation of isoflavones on their bioavailability and pharmacokinetics in aged male rats. Mol. Nutr. Food Res. 2009, 53 (Suppl. 1), S16–S26. (11) Chang, Y. C.; Nair, M. G.; Santell, R. C.; Helferich, W. G. Microwave-mediated synthesis of anticarcinogenic isoflavones from soybeans. J. Agric. Food Chem. 1994, 42 (9), 1869–1871. (12) Nurmi, T.; Mazur, W.; Heinonen, S.; Kokkonen, J.; Adlercreutz, H. Isoflavone content of the soy based supplements. J. Pharm. Biomed. Anal. 2002, 28 (1), 1–11. (13) Wu, S.; Murphy, P. A.; Johnson, L. A.; Reuber, M. A.; Fratzke, A. R. Simplified process for soybean glycinin and beta-conglycinin fractionation. J. Agric. Food Chem. 2000, 48 (7), 2702–2708. (14) Penalvo, J. L.; Nurmi, T.; Adlercreutz, H. A simplified HPLC method for total isoflavones in soy products. Food Chem. 2004, 87 (2), 297–305. (15) Reeves, P. G.; Nielsen, F. H.; Fahey, G. C., Jr. AIN-93 purified diets for laboratory rodents: final report of the American Institute of Nutrition ad hoc writing committee on the reformulation of the AIN-76A rodent diet. J. Nutr. 1993, 123 (11), 1939–1951. (16) Nutrient Requirements of Laboratory Animals, 4th revised ed.; National Academy Press: Washington, DC, 1995; p 192. (17) U.S. Department of Agriculture, Agricultural Research Service. USDA National Nutrient Database for Standard Reference. Release 22; Nutrient Data Laboratory Home Page, 2009; http://www.ars. usda.gov/nutrientdata. (18) Twaddle, N. C.; Churchwell, M. I.; Doerge, D. R. High-throughput quantification of soy isoflavones in human and rodent blood using liquid chromatography with electrospray mass spectrometry and tandem mass spectrometry detection. J. Chromatogr., B: Anal. Technol. Biomed. Life Sci. 2002, 777 (1-2), 139–145. (19) Fukutake, M.; Takahashi, M.; Ishida, K.; Kawamura, H.; Sugimura, T.; Wakabayashi, K. Quantification of genistein and genistin in soybeans and soybean products. Food Chem. Toxicol. 1996, 34 (5), 457–461. (20) Doerge, D. R.; Chang, H. C.; Churchwell, M. I.; Holder, C. L. Analysis of soy isoflavone conjugation in vitro and in human blood using liquid chromatography-mass spectrometry. Drug. Metab. Dispos. 2000, 28 (3), 298–307. (21) Sepehr, E.; Cooke, G.; Robertson, P.; Gilani, G. S. Bioavailability of soy isoflavones in rats Part I: application of accurate methodology for studying the effects of gender and source of isoflavones. Mol. Nutr. Food Res. 2007, 51 (7), 799–812. (22) Setchell, K. D.; Zimmer-Nechemias, L.; Cai, J.; Heubi, J. E. Exposure of infants to phyto-oestrogens from soy-based infant formula. Lancet 1997, 350 (9070), 23–27. (23) Castellanos, V. H.; Litchford, M. D.; Campbell, W. W. Modular protein supplements and their application to long-term care. Nutr. Clin. Pract. 2006, 21 (5), 485–504. (24) Eldridge, A. C.; Kwolek, W. F. Soybean isoflavones: Effect of environment and variety on composition. J. Agric. Food Chem. 1983, 31 (2), 394–396. (25) Seguin, P.; Zheng, W.; Smith, D. L.; Deng, W. Isoflavone content of soybean cultivars grown in eastern Canada. J. Sci. Food Agric. 2004, 84 (11), 1327–1332.

4536

J. Agric. Food Chem., Vol. 58, No. 7, 2010

(26) Kim, S. H.; Jung, W. S.; Ahn, J. K.; Kim, J. A.; Chung, I. M. Quantitative analysis of the isoflavone content and biological growth of soybean (Glycine max L.) at elevated temperature, CO2 level and N application. J. Sci. Food Agric. 2005, 85 (15), 2557– 2566. (27) Rasolohery, C. A.; Berger, M.; Lygin, A.; Lozovaya, V.; Nelson, R. L.; Dayde, J. Effect of temperature and water availability during late maturation of the soybean seed on germ and cotyledon isoflavone content and composition. J. Sci. Food Agric. 2008, 88 (2), 218–228. (28) Aguiar, C. L.; Baptista, A. S.; Walder, J. M.; Tsai, S. M.; CarraoPanizzi, M. C.; Kitajima, E. W. Changes in isoflavone profiles of soybean treated with gamma irradiation. Int. J. Food Sci. Nutr. 2009, 1–8. (29) Genovese, M. I.; Barbosa, A. C.; Pinto Mda, S.; Lajolo, F. M. Commercial soy protein ingredients as isoflavone sources for functional foods. Plant Foods Hum. Nutr. 2007, 62 (2), 53–58. (30) Wang, C.; Ma, Q.; Pagadala, S.; Sherrard, M. S.; Krishman, P. G. Changes of isoflavones during processing of soy protein isolates. J. Am. Oil Chem. Soc. 1998, 75 (3), 337–341. (31) Yin, L. J.; Li, L. T.; Liu, H.; Saito, M.; Tatsumi, E. Effects of fermentation temperature on the content and composition of isoflavones and beta-glucosidase activity in sufu. Biosci., Biotechnol., Biochem. 2005, 69 (2), 267–272. (32) Huang, R. Y.; Chou, C. C. Stability of isoflavone isomers in steamed black soybeans and black soybean Koji stored under different conditions. J. Agric. Food Chem. 2009, 57 (5), 1927–1932. (33) Lee, S. J.; Ahn, J. K.; Kim, S. H.; Kim, J. T.; Han, S. J.; Jung, M. Y.; Chung, I. M. Variation in isoflavone of soybean cultivars with location and storage duration. J. Agric. Food Chem. 2003, 51 (11), 3382–3389. (34) Lin, J.; Krishnan, P. G.; Wang, C. Retention of isoflavones and saponins during the processing of soy protein isolates. J. Am. Oil Chem. Soc. 2006, 83 (1), 59–63. (35) Setchell, K. D.; Cole, S. J. Variations in isoflavone levels in soy foods and soy protein isolates and issues related to isoflavone databases and food labeling. J. Agric. Food Chem. 2003, 51 (14), 4146–4155. (36) USDA-Iowa State University, Agricultural Research Service. Database on the Isoflavone Content of Foods; Nutrient Data Laboratory Website, 2002; http://www.nal.usda.gov/fnic/foodcomp/Data/ isoflav/isoflav.html. (37) Zhou, S.; Hu, Y.; Zhang, B.; Teng, Z.; Gan, H.; Yang, Z.; Wang, Q.; Huan, M.; Mei, Q. Dose-dependent absorption, metabolism, and excretion of genistein in rats. J. Agric. Food Chem. 2008, 56 (18), 8354–8359. (38) Qiu, F.; Chen, X. Y.; Song, B.; Zhong, D. F.; Liu, C. X. Influence of dosage forms on pharmacokinetics of daidzein and its main metabolite daidzein-7-O-glucuronide in rats. Acta Pharmacol. Sin. 2005, 26 (9), 1145–1152. (39) Kwon, S. H.; Kang, M. J.; Huh, J. S.; Ha, K. W.; Lee, J. R.; Lee, S. K.; Lee, B. S.; Han, I. H.; Lee, M. S.; Lee, M. W.; Lee, J.; Choi, Y. W. Comparison of oral bioavailability of genistein and genistin in rats. Int. J. Pharm. 2007, 337 (1-2), 148–154. (40) Jefferson, W. N.; Doerge, D. R.; Padilla-Banks, E.; Woodling, K. A.; Kissling, G. E.; Newbold, R. R. Oral exposure to genistin, the glycosylated form of genistein, during neonatal life adversely affects the female reproductive system. Environ. Health Perspect. 2009, 117, 1883–1889.

Andrade et al. (41) Holder, C. L.; Churchwell, M. I.; Doerge, D. R. Quantification of soy isoflavones, genistein and daidzein, and conjugates in rat blood using LC/ES-MS. J. Agric. Food Chem. 1999, 47 (9), 3764–3770. (42) McClain, R. M.; Wolz, E.; Davidovich, A.; Pfannkuch, F.; Edwards, J. A.; Bausch, J. Acute, subchronic and chronic safety studies with genistein in rats. Food Chem. Toxicol. 2006, 44 (1), 56–80. (43) Izumi, T.; Piskula, M. K.; Osawa, S.; Obata, A.; Tobe, K.; Saito, M.; Kataoka, S.; Kubota, Y.; Kikuchi, M. Soy isoflavone aglycones are absorbed faster and in higher amounts than their glucosides in humans. J. Nutr. 2000, 130 (7), 1695–1699. (44) Setchell, K. D.; Brown, N. M.; Desai, P.; Zimmer-Nechemias, L.; Wolfe, B. E.; Brashear, W. T.; Kirschner, A. S.; Cassidy, A.; Heubi, J. E. Bioavailability of pure isoflavones in healthy humans and analysis of commercial soy isoflavone supplements. J. Nutr. 2001, 131 (4 Suppl.), 1362S–1375S. (45) Richelle, M.; Pridmore-Merten, S.; Bodenstab, S.; Enslen, M.; Offord, E. A. Hydrolysis of isoflavone glycosides to aglycones by beta-glycosidase does not alter plasma and urine isoflavone pharmacokinetics in postmenopausal women. J. Nutr. 2002, 132 (9), 2587–2592. (46) Rufer, C. E.; Bub, A.; Moseneder, J.; Winterhalter, P.; Sturtz, M.; Kulling, S. E. Pharmacokinetics of the soybean isoflavone daidzein in its aglycone and glucoside form: a randomized, double-blind, crossover study. Am. J. Clin. Nutr. 2008, 87 (5), 1314–1323. (47) Nielsen, I. L.; Williamson, G. Review of the factors affecting bioavailability of soy isoflavones in humans. Nutr. Cancer 2007, 57 (1), 1–10. (48) Busby, M. G.; Jeffcoat, A. R.; Bloedon, L. T.; Koch, M. A.; Black, T.; Dix, K. J.; Heizer, W. D.; Thomas, B. F.; Hill, J. M.; Crowell, J. A.; Zeisel, S. H. Clinical characteristics and pharmacokinetics of purified soy isoflavones: Single-dose administration to healthy men. Am. J. Clin. Nutr. 2002, 75 (1), 126–136. (49) King, R. A.; Broadbent, J. L.; Head, R. J. Absorption and excretion of the soy isoflavone genistein in rats. J. Nutr. 1996, 126 (1), 176–182. (50) Zubik, L.; Meydani, M. Bioavailability of soybean isoflavones from aglycone and glucoside forms in American women. Am. J. Clin. Nutr. 2003, 77 (6), 1459–1465. (51) Cassidy, A.; Brown, J. E.; Hawdon, A.; Faughnan, M. S.; King, L. J.; Millward, J.; Zimmer-Nechemias, L.; Wolfe, B.; Setchell, K. D. Factors affecting the bioavailability of soy isoflavones in humans after ingestion of physiologically relevant levels from different soy foods. J. Nutr. 2006, 136 (1), 45–51. (52) Muthyala, R. S.; Ju, Y. H.; Sheng, S.; Williams, L. D.; Doerge, D. R.; Katzenellenbogen, B. S.; Helferich, W. G.; Katzenellenbogen, J. A. Equol, a natural estrogenic metabolite from soy isoflavones: convenient preparation and resolution of R- and S-equols and their differing binding and biological activity through estrogen receptors alpha and beta. Bioorg. Med. Chem. 2004, 12 (6), 1559–1567. Received for review November 12, 2009. Revised manuscript received January 4, 2010. Accepted February 17, 2010. This research was funded by the National Institute on Aging with additional support from National Institute for Complementary and Alternative Medicine, Office of Dietary Supplements, and Women’s Health Initiative (P01 AG024387 to W.G.H. and D.R.D.). The views presented in this article do not necessarily reflect those of the U.S. Food and Drug Administration.